VOLUME 75 FALL 2016

STRATIGRAPHIC NOMENCLATURE OF THE OUACHITA IN : FORMAL, INFORMAL, OBSOLETE, AND INCORRECT or THE GOOD, THE BAD, AND THE UGLY INSIDE ON PAGE 5 OKLAHOMA GEOLOGICAL SURVEY

OKLAHOMA GEOLOGICAL SURVEY

Dr. Jeremy Boak, Director

OGS Geologist Neil Suneson Editor delves into the Good, the Bad, and Ted Satterfield the Ugly of the topic of stratigraphic nomenclature in the .

Cartography Manager James Anderson

GIS Specialist Russell Standridge

Copy Center Manager Richard Murray

This publication, printed by the Oklahoma Geological Survey, Norman, Oklahoma, is issued by the Oklahoma Geological Survey as authorized by Title 70, Oklahoma Statutes 1981, Section 3310, and Title 74, Oklahoma Statutes 1981, Sections 231—238

The Oklahoma Geological Survey is a state agency for research and public service, mandated in the State Constitution to study Oklahoma’s land, water, mineral and energy resources and to promote wise Cover: Outcrop of the Wapanucka Limestone use and sound environmental practices. east of Hartshorne, Oklahoma. (Photo by TED SATTERFIELD). 2 FALL 2016

Grateful to be back

As the Director of the Oklahoma Geological Survey (OGS), I am very glad to see the revival of the Oklahoma Geology Notes, which had been a feature of the Oklahoma geology scene since the early 40s, but was last published in 2014. This issue, marking the return, introduces a new format, becoming less of a newsletter and more of a scientific publication. From there we proceed, as many geological investigations do, from the surface downward. We start up the Notes again with Neil Sune- son’s take on stratigraphic nomenclature in the Ouachita Mountains. Just the word “nomen- clature” tends to produce yawns among many scientists, but making sure that we are talking about the same rock units can be a challenge when different regional names don’t necessarily meet in the middle due to the lateral variations Jeremy Boak of rock units. OGS Director As geologists, we are tasked with character- izing the crust of the earth in three spatial di- Providing clear rationale, and straightforward mensions, through time. But, as OGS geophysi- description of any revision of rock names — cist Kevin Crain pointed out to me recently, the what names are being changed, and how do the problem is really n-dimensional, as we must new names relate to existing stratigraphic terms identify and measure critical properties, each — is needed to ensure that the value of older of which covers its own natural range, varying literature will not be lost. through time as well. In the realm of sedimen- Our hope is that the new Oklahoma Geology tary rocks, we can count in general on greater Notes will channel the best of the previous long variation in the vertical than the horizontal series and bring some new scenes, new takes on dimension (or we would not be able to iden- old locations, current developments, interesting tify strata to begin with). But the recognition ideas, and show that, no matter how old, there that lithostratigraphic units are not the same as is always something new to be learned about chronostratigraphic units can lead to confusion. Oklahoma geology.

3 OKLAHOMA GEOLOGICAL SURVEY

4 FALL 2016

STRATIGRAPHIC NOMENCLATURE OF THE OUACHITA MOUNTAINS IN OKLAHOMA: FORMAL, INFORMAL, OBSOLETE, AND INCORRECT or The Good, the Bad, and the Ugly By Neil H. Suneson

INTRODUCTION units that are exposed or are in the subsurface in the Oklahoma Ouachita Mountains and that are present The nomenclature of the stratigraphic units in the subsurface near the Choctaw Fault. The units (groups, formations, members) in the Ouachita Moun- in the subsurface in the southern part of the Arkoma tains in Oklahoma has changed since the first geologic Basin are typically folded and faulted and are part of map of part of the range was published in the Ouachita tectonic belt. The names are limited to 1902 (Taff, 1902). Proposed revisions to the formation those of units that are older than the Desmoinesian names were typically based on a number of factors: Hartshorne Formation. Whereas the Hartshorne and 1) more detailed mapping than was previously avail- younger formations are folded and faulted, they are not able; 2) more and better paleontological data; 3) bet- present at the surface south of the trace of the Choctaw ter stratigraphic correlations with units exposed in the Fault and are therefore beyond the scope of this paper. Oklahoma and Ouachita Mountains and the Arkoma Basin; and 4) more and better subsurface data. BACKGROUND In addition, changes in the “rules” for naming geologic units as prescribed in the 1933, 1961, 1970, 1983, and The basis for this paper is the Code of Stratigraphic 2005 “Codes of Stratigraphic Nomenclature” forced Nomenclature. As discussed below, many names were Ouachita stratigraphers and geologists to reexamine assigned before any code existed. Ashley et al. (1933) formerly accepted names. In many cases, however, published the first “code” (although the term “code” geologists ignored the existing code and introduced was not included in the title), and two of its co-authors new terms without emphasizing their informal nature. – Charles Gould and Hugh Miser – were very familiar Petroleum geologists exacerbated the confusion, par- with the stratigraphy of the Ouachita Mountains. The ticularly in the northern frontal belt (Figure 1) of the authors used the word “rules” in their paper, but ad- Ouachitas, by using new names that applied to strata mitted that the “rules” were really “guidelines.” The encountered only in the subsurface. American Commission on Stratigraphic Nomenclature This report briefly describes the history of the (ACSN) published a revised code in 1961 (ACSN, stratigraphic nomenclature of the Ouachita Moun- 1961) and an only slightly revised code again in 1970 tains in Oklahoma. It describes the currently accepted (ACSN, 1970). The code was largely rewritten in 1983 group, formation (Figure 2), and member names and by the North American Commission on Stratigraphic distinguishes between those that are generally con- Nomenclature (NACSN) (NACSN, 1983) who also sidered “formal” by geologists and those that, while published a revised version in 2005 (NACSN, 2005). widely used, are informal. This report also comments Like the 1933 “code,” the 2005 version is more a set on some problems with the accepted formal and in- of guidelines than rules as evidenced by the common formal nomenclature and proposes some solutions to use of the word “should” instead of “must.” For exam- those problems. The subsurface nomenclature used ple, when naming a subsurface unit, “the hole or mine in the northern frontal belt is discussed and clarified. should be (instead of “must be”) located precisely” More importantly, perhaps, this paper recommends that (NACSN, 2005, p. 1564) (author’s italics and paren- some commonly used but informal names be viewed theses). as formal and that formal, but little-used or poorly jus- Many of the units in the Oklahoma Ouachita tified, names or those of local units be changed. Mountains were named before the 1933 “code” was The names reviewed in this report are of those published and have been widely used for almost 100

5 OKLAHOMA GEOLOGICAL SURVEY

Figure 1. Generalized geologic map of the Ouachita Mountains, Oklahoma. Modified from Arbenz (2008, plate 2). years. Although their names do not fulfil many of the logic Names Lexicon (“Geolex”) (http://ngmdb.usgs. criteria now necessary for a name to be recognized, gov/Geolex/search). Geolex is a searchable database they are considered formal because they have been based on lexicons published as hard-copy bulletins by used for many years and are well-established in the the USGS. This database lists names used by the USGS geologic literature. Article 7 (c) of the modern strati- and the different state surveys and claims to identify graphic code (NACSN, 2005) recognizes the need with a slash (see website) those that do not conform to preserve commonly used, but perhaps imprecisely with the stratigraphic code(s) cited above; in fact, defined, names of stratigraphic units. One unit – the Geolex does not do this (e.g., see discussion on Lynn Spiro Sandstone at the base of the Atoka Formation – Mountain Formation, below). Geolex also does not al- is widely used in the petroleum industry and is easily ways restrict the name of a unit to a specific rank; for recognized and mapped on the surface and subsurface. example, the USGS and OGS recognize the Springer Although it has not met the rigorous requirements for as a group and a formation. In addition, the OGS rec- being considered formal, this paper recognizes it as ognizes Springer Shale as a formal unit. Table 1 shows such. the names of the units in the Oklahoma Ouachita An important source for stratigraphic nomencla- Mountains as recognized by the USGS and Oklahoma ture is the U.S. Geological Survey’s (USGS) U.S. Geo- Geological Survey (OGS) (source: Geolex database at

6 STRATIGRAPHIC COLUMN OF OUACHITA MOUNTAINS

FALL 2016

Platform Stratigraphy Ouachita Stratigraphy n a i Atoka Formation Atoka Formation n Atokan a v l y

s Wapanucka Limestone Johns Valley Shale n n e Jackfork Group P

Morrowan Springer Formation Chest- n erian a

i Stanley Group

p Mera- p

i mecian s

s Caney Shale i

s Osagean s i

M Kinder- hookian Arkansas Novaculite Upper Middle

Devonian Lower Upper Lower Silurian Polk Creek Shale Upper

Middle

Ordovician Lower Crystal Mountain Sandstone

Upper Collier Shale Cambrian

Figure 2. Generalized stratigraphic column of the Ouachita Mountains, Oklahoma. Subdivisions of groups and formations discussed in text. Modified from Arbenz (2008, plate 2).

7 OKLAHOMA GEOLOGICAL SURVEY ngmdb.usgs.gov/Geolex/search; accessed June 28, 2016). In addition, proposed changes to and problems with generally accepted nomenclature are shown; the reasons for the changes are sum- marized in the table and/or discussed in the text. Throughout this report, unit names with all capital first letters are considered or are- recom mended to be formal in this report.

FORMAL STRATIGRAPHY OF THE OUACHITA MOUNTAINS, OKLAHOMA

INTRODUCTION Figure 3. Atoka Formation. Location: on south side of US Highway The generally accepted stratigraphy of the 259 just south of intersection with US Highway 59. SE NW SE sec. 17, Ouachita Mountains in Oklahoma is shown in T. 3 N., R. 26 E. Photograph taken in 1988 and is now more vegetated. Figure 2. The names and ranks of all of the units shown are recognized by the USGS and are used by the OGS. The following section briefly the Atoka Formation. Two publications (Pitt et al., describes the origin of the names used by geologists 1982; Marcher and Bergman, 1983) refer to the Atoka mapping in southeastern Oklahoma; most of them Formation as the Lynn Mountain Formation; this name were established before any code of stratigraphic no- is discussed below. In addition, the petroleum industry menclature existed and are accepted because they have recognizes several named producing units in the Atoka been referred to in the geological literature for nearly Formation; these are also discussed below. 100 years. Some of the formally named formations, however, should be reexamined; in particular, the De- Spiro Sandstone Member, Atoka Formation vonian formations in the frontal belt (Figure 1) that may be olistoliths in the Johns Valley Shale and the The Spiro sandstone member (of the Desmoinesian formations that make up the Jackfork Group. Savanna Formation, Krebs Group) was used by Wilson (1935) and Wilson and Newell (1937) for exposures in FRONTAL BELT Muskogee and McIntosh Counties but was named for sandstone outcrops near the town of Spiro in Le Flore Atoka Formation (Atokan) County. Wilson (1935) believed the Spiro sandstone in Muskogee County was the same as a sandstone that The youngest formation exposed in the Ouachita was mapped (unpublished) near Spiro by W.T. Thom, Mountains and the oldest exposed in the Arkoma Ba- Jr. Wilson (1935) did not formally designate a type sec- sin is the Atoka Formation (Figure 3). The name first tion but did describe, in a very general way, the Spiro appeared in a report (Taff and Adams, 1900) on the sandstone. Knechtel (1949) did not recognize Wilson’s eastern Choctaw coal fields in what was then Indian Spiro sandstone in northern Le Flore County and Jor- Territory. The coal fields, however, did not extend into dan (1957) noted that the OGS abandoned the name either Atoka County or the town of Atoka. Taff (1902) as a member of the Savanna Formation. Geolex, how- later mapped the formation in the Atoka County area, ever, erroneously shows this Spiro sandstone as being but never designated a type section. The name “Atoka used by the OGS. Formation” has been used in all (but two) publications An identically named Spiro Sandstone (or “sand”) on the geology of the Ouachita Mountains and Arkoma (of the Atoka Formation) (Figure 4) is a prolific gas Basin, most notably Hendricks (1939), Hendricks et al. reservoir in the southern part of the Arkoma Basin and (1947), Cline (1960), Shelburne (1960), Seely (1963), frontal belt of the Ouachita Mountains (Lumsden et al., Hart (1963), and Fellows (1964). Geologic maps pro- 1971; Grayson and Hinde, 1993; Gross et al., 1995) and duced by the OGS as part of the COGEOMAP and is widely recognized by petroleum geologists working STATEMAP programs beginning in 1989 also show in the area. (It is also referred to in well completion

8 FALL 2016

Figure 4. Spiro Sandstone Member, Atoka Formation. Location: along Pittsburg – Latimer County line about 2.5 mi east-south- east of Hartshorne and just south of Oklahoma Highway 1/63. SW SW SW sec. 10, T. 4 N., R. 17 E. Photograph taken by Ted Satterfield, OGS. reports as “basal Atoka” or “lowermost Atoka” sand- ma, that it is considered a formal member of the Atoka stone.) It was first named in the Le Flore Gas and Elec- Formation in this report and is therefore capitalized, tric Company No. 1 Frank Parnell well (NE SE sec. despite not being listed in the USGS Geolex database. 18, T. 9 N., R. 27 E.), located about eight miles east of the town of Spiro, and was the discovery well for the Wapanucka Limestone (formation) (Morrowan) Cedars Southwest (now Cedars) gas field (Maravich, 1955). This Spiro Sandstone also crops out through- The Wapanucka Limestone (Figure 5) underlies out the central part of the frontal belt of the Ouachi- the Atoka Formation and overlies the Springer For- ta Mountains where it underlies turbidite sandstones mation in the Ouachita Mountains frontal belt. It was and shales that form most of the Atoka Formation and named by Taff (1901) probably for exposures near the overlies the Wapanucka Limestone or the Johns Valley town of Wapanucka in Johnston County, but he did Shale (OGS COGEOMAP and STATEMAP geologic not specify a type locality or type section. The Wapa- maps). nucka Limestone appears on detailed geologic maps The code of stratigraphic nomenclature allows of the Ouachita Mountains frontal belt, including Hen- subsurface units to be formalized (Article 16), but the dricks et al. (1947) and the northwestern COGEMAP Spiro Sandstone (Atoka Formation) has not met sev- and STATEMAP maps published by the OGS. Harlton eral of the criteria for formal recognition. However, the (1938) proposed subdividing Taff’s Wapanucka into name “Spiro Sandstone” is now so established in the three formations, from oldest to youngest, the Prim- geologic literature, so widely recognized and mappa- rose Formation, the Limestone Gap Shale, and the ble in outcrop, so easily recognized on well logs, and Wapanucka Limestone; however, despite Harlton’s so important as a gas reservoir in this part of Oklaho- detailed stratigraphic work, this proposal was not fol-

9 OKLAHOMA GEOLOGICAL SURVEY lowed by subsequent workers. Springer Formation (Morrowan and Chesterian) Chickachoc Chert Member, Wapanucka Limestone (formation) The Springer Formation is exposed not only in the Ouachita Mountains where it underlies the Wapa- Taff (1901) named the Chickachoc Chert in his nucka Limestone and overlies the Caney Shale, but in study of the western part of the Arkoma Basin, pre- the and Ardmore Basin where it sumably for the Chickiechockie Post Office (current has been far more thoroughly studied. The Springer name Chockie) located just east of his map area. Like is also referred to as the Springer Shale and locally all of the units named by Taff, neither a type area, a is recognized as a group where it is thick, paleonto- type locality, nor a type section was identified. Taff logically well dated, and contains numerous sandstone (1901) believed the chert was a lentil within the Atoka beds. Goldston (1922) named the unit for the town of Formation and Hendricks et al. (1947) mapped it as Springer in the Ardmore Basin and recognized it as a a formation separate from, but correlative with, the member of the Glenn Formation, but he did not iden- Wapanucka Limestone. Grayson (1979) identified the tify a type section. Tomlinson (1929) raised its rank Chickachoc Chert as the lower informal member of the from member to formation based partly on the pres- Wapanucka Formation present only in the more south- ence of persistent mappable sandstone beds that he eastern thrust sheets of the Ouachita Mountains. The considered as members, but did not identify a type sec- USGS recognizes Chickachoc Chert and Chickachoc tion or locality. Chert Member despite its uncommon usage and lack Despite its lack of detailed study, the Springer of formal publication. in the Ouachita Mountains is considered a formation

Figure 5. Wapanucka Limestone. Location: along Oklahoma Highway 1/63 three miles east-southeast of Hartshorne. NE SE SW sec. 10, T. 4 N., R. 17E. Photograph taken by Rick Andrews, OGS.

10 FALL 2016 by the USGS and OGS. It is present in several thrust Arbuckle Mountains (Elias, 1956; Elias and Branson, sheets in the western part of the Ouachita Mountains 1959).) Modern geologic maps (especially Hendricks frontal belt (Figure 1) where it is underlain by the et al., 1947) show no Caney Shale in Johns Valley; the Caney Shale (Hendricks et al., 1947) and is present as Caney outcrops there are allochthonous submarine far east as the town of Red Oak (Hemish et al., 1990). slide masses (olistostromes) within the Johns Valley Shale. Thus, the Caney Shale in its presumed type area Caney Shale (formation) (Chesterian to Kinder- is stratigraphically out-of-place. The failure to recog- hookian) nize the allochthonous nature of the Caney Shale in the Johns Valley Shale led many geologists to suggest that The Caney Shale (Figure 6) has one of the more in- the underlying Jackfork Group was Mississippian. teresting nomenclatural histories in southern Oklaho- Hendricks et al. (1947) mapped what they believed ma. Much of the discussion focusses on the age of the to be stratigraphically autochthonous Caney Shale in Caney (both “Mississippian Caney” and “Pennsylva- the western part of the Ouachita Mountains frontal nian Caney” appear in numerous papers) and its rela- belt (Figure 1). Most of these outcrops(?) are shown tion to the overlying shale units; most of these discus- as long narrow bands at the base of thrust sheets over- sions are based on exposures in the Ardmore Basin and lain by even narrower bands of the Springer Forma- adjacent Arbuckle Mountains and have little bearing tion. Very few strike and dip measurements are shown on how the Caney is used in the Ouachita Mountains along these bands; therefore, their true continuity is which is where the Caney was named. Ulrich (1927) unknown. Because these Caney Shale outcrops are au- was the first to describe the differences between what tochthonous, the Caney Shale is included on Figure 2. he called the “Ouachita Caney” and the “Arbuckle However, other, typically smaller areas of the Caney Caney.” The result of many of these studies was that Shale are underlain by the Woodford Shale and, in one the name “Caney” and origin of the formation were the case, the Woodford Shale is underlain by the Pinetop source of years of misinterpreted geology. Chert. Suneson and Ferguson (1990) mapped a Wood- Taff (1901) first named the unit in his work on the ford – Caney outcrop in the eastern part of the frontal Coalgate quadrangle which is mostly in the Arkoma belt as a large olistolith in the Johns Valley Shale. The Basin, but he probably named it after Cane Creek (Pinetop) – Woodford – Caney outcrops mapped by (Caney Creek on some maps) in the Tuskahoma Syn- Hendricks et al. (1947) are probably also slide masses cline about 20 mi to the east (Cline, 1960, p. 60). Cane within the Johns Valley (Suneson and Hemish, 1994a, Creek is one of several streams that drain Johns Valley, p. 74-75); therefore, these units are not included in a the type locality of the Johns Valley Shale (discussed stratigraphic column of the Ouachita Mountains (Fig- below). (Additional confusion regarding the name ure 2). “Caney” is described by Elias and Branson (1959, p. 4), who also proposed a type section for the Caney in the Woodford Shale (formation) (Chesterian to Upper Devonian)

The Woodford Shale (also Chert) was named by Taff (1902) in his work on the Atoka quad- rangle, presumably for outcrops near the town of Woodford in Carter County about 50 mi to the west. Because many, if not all, of the Woodford outcrops in the Ouachita Mountains are prob- ably stratigraphically allochthonous and are con- tained within the Johns Valley Shale, the unit is not described in this report.

Pinetop Chert (formation) (Lower Devonian)

Miser (1934, p. 974) named the Pinetop Chert for exposures near Pinetop School in the Figure 6. Caney Shale. Location: along Brushy Creek. NW SW NW Ouachita Mountains and Hendricks et al. (1947) sec. 5, T. 2 N., R. 15 E. showed the single, small outcrop of the unit on

11 OKLAHOMA GEOLOGICAL SURVEY their map. Amsden (1983) correlated the Pi- netop with the Haragan and Bois d’Arc Lime- stones of the Hunton Group and Ham (1959) Table I. — Formal Nomenclature correlated it with the lower part of the Arkansas of the Ouachita Mountains, Oklahoma Novaculite. However, as noted above in the de- scription of the Woodford Shale and by Sune- Frontal Belt son and Campbell (1990), this unit probably is stratigraphically allochthonous and therefore is Atoka Formation not described in this report. Spiro Sandstone Member (1) Wapanucka Limestone (formation) CENTRAL BELT Chickachoc Chert Member Springer Formation Atoka Formation (Atokan) Caney Shale (formation) (3)

The Atoka Formation (Figure 3) in the Central Belt Ouachita Mountains central belt (Figure 1) dif- fers little from that in the frontal belt and is de- Atoka Formation scribed above. Johns Valley Shale (formation) Jackfork Group Johns Valley Shale (formation) (Morrowan) Game Refuge Sandstone (formation) Wesley Shale (formation) (3) The nomenclatural history of the Johns Val- Markham Mill Formation (3) ley Shale is closely related to that of the Caney Prairie Mountain Formation (3) Shale described above. Ulrich (1927) proposed Prairie Hollow Shale Member (2) the name Johns Valley Shale for the same unit Wildhorse Mountain Formation Taff (1901) called Caney Shale, and Cane Creek Stanley Group (also called Caney Creek), Taff’s (1901) pre- Chickasaw Creek Shale (formation) sumed type locality of the Caney Shale, is one Moyers Formation of several creeks that drain the bowl-shaped de- Tenmile Creek Formation pression known as Johns Valley. Although Ul- rich (1927) identified Johns Valley as the type Broken Bow Uplift locality of the Johns Valley Shale, he did not identify a type section, probably because the Stanley Group (Chesterian to Osagean) exposures in the area are very poor. Tenmile Creek Formation The Johns Valley Shale is widely recognized Hatton Tuff Lentil throughout the Ouachita Mountains (Hendricks et al., 1947; Cline, 1960; Shelburne, 1960; Seely, Arkansas Novaculite (formation) 1963; Hart, 1963; Fellows, 1964; Pitt et al., Missouri Mountain Shale (formation) 1982; and OGS COGEOMAP and STATEMAP Blaylock Sandstone (formation) geologic maps) although its distribution is not Polk Creek Shale (formation) agreed upon. Early workers (e.g., Miser, 1929, Bigfork Chert (formation) p. 25; Cline, 1960, p. 17; Hendricks et al., 1947) Womble Shale (formation) restricted exposures of the Johns Valley to south Blakely Sandstone (formation) of the Ti Valley Fault and suggested that the Mazarn Shale (formation) pre-Atoka Pennsylvanian and Mississippian Crystal Mountain Sandstone (formation) formations north of the fault resembled those Collier Shale (formation) in the Arkoma Basin whereas those units south of the fault consisted of deep-water turbidites. Footnotes: no number – unit recognized by USGS More recent mapping by the OGS in the frontal and OGS; (1) recommended addition of unit (see belt has shown that the Johns Valley is present text). (2) recommended change of unit (see text). north of the Ti Valley Fault (Suneson, 1988). (3) Caney Shale (of Springer Group) used by This difference is partly related to the interpret- USGS; Wesley, Markham Mill, and Prairie Hollow require re-evaluation (see text)

12 FALL 2016 ed origins of the (Pinetop) – Woodford – Caney out- (1977) who showed that the immediately underlying crops discussed above. Are the large, stratigraphically Chickasaw Creek Shale was uppermost Chesterian coherent blocks slide masses (large olistoliths) within and that most of the Jackfork was Morrowan, thereby the Johns Valley Shale, in which case the Johns Valley agreeing with Hendricks et al. (1947) and most previ- is present north of the Ti Valley Fault in the western ous workers. Whiteside and Grayson (1990) confirmed Ouachitas, or do they form the base of thrust sheets? the Pennsylvanian age using . The origin of these exposures, as well as the origin of Harlton (1938) elevated the rank of the Jackfork to the Johns Valley Shale itself, is beyond the scope of group status and divided it into four formations – from this report, but both are critical for determining how to bottom to top, the Wildhorse Mountain, Prairie Moun- map and interpret the formation. tain (including the Prairie Hollow Member), Markham Mill, and Wesley. Harlton (1938) based his divisions Jackfork Group (Morrowan) on what he believed to be widespread siliceous shales, although it is unclear what his evidence for this is be- Taff (1902) first named the Jackfork sandstone in cause his published geologic maps are of relatively his work on the Atoka quadrangle for Jackfork Moun- small areas. (As described below, many later work- tain, which is about 15 mi east of his map area. Taff ers failed to find Harlton’s siliceous shales.) In 1959 (1902, p. 4) did not designate a type section; he merely Harlton added a fifth, uppermost formation — the stated, “it receives its name from Jackfork Mountain, Game Refuge — and moved the Prairie Hollow Shale in which a large part of the formation is exposed.” Member into the Wildhorse Mountain (Figure 8). The Much later Pitt et al. (1982, p. 80) attempted to iden- nomenclatural history of these are described below. tify a type section on Jackfork Mountain but ignored The Jackfork Group names used in Oklahoma are not many of the recommendations set forth in the then- recognized in Arkansas, rather, the Jackfork is divided accepted Code of Stratigraphic Nomenclature (ACSN, into an upper Brushy Knob Formation and lower Irons 1970). Their lithologic descriptions are brief and much Fork Mountain Formation (Morris, 1971). of the unit, including the top and base, is covered. In Harlton’s (1938) formation names were used by addition, part of their “type section” crosses a fault and many subsequent workers (e.g., Cline, 1960; Shelburne, includes part of the Stanley Group (Suneson, 1991). 1960; Hart,1963; Seely, 1963; Fellows, 1964, Briggs, As a result the Jackfork, like most of the Mississippian 1973) in the central and eastern Oklahoma Ouachita and Pennsylvanian units in the Ouachita Mountains, Mountains, but many of the formations were combined has a type area or type locality but does not have a type (shown as “undivided” on the geologic maps) because section. the siliceous shales used by Harlton (1938) were “ob- The Jackfork Group (formerly sandstone or forma- scure,” unmappable, or absent (Figure 8). The only tion) (Figure 7) is widely recognized as a formal unit units within the Jackfork Group that were mapped and/ in the Ouachita Mountains and is shown on most of the or recognized by most workers are the Game Refuge detailed maps of the region (Honess, 1923; Hendricks Sandstone (formation) and the Prairie Hollow Shale et al., 1947; Cline, 1960; Shelburne, 1960; Seely, Member, and most workers followed Harlton (1959) 1963; Hart, 1963; Fellows, 1964; Briggs, 1973; Pitt and Cline (1960) who included the Prairie Hollow in et al., 1982; and OGS COGEOMAP and STATEMAP the Wildhorse Mountain Formation and not the Prairie geologic maps). The age of the Jackfork, however, has Mountain Formation (Harlton, 1938). Hendricks et al. been controversial mostly because the age (and origin) (1947) recognized the widespread distribution of the of the overlying Johns Valley Shale was controversial. siliceous shales in the Jackfork Sandstone (his term) Cline (1956) first proposed that the Jackfork was Mis- in the western Ouachita Mountains but did not use the sissippian based entirely on his belief that the base of formation names suggested by Harlton (1938). Morris the overlying Johns Valley consisted of in situ “black (1971) was not able to map Harlton’s (1938) siliceous shale of typical Caney lithology” (p. 105). Cline (1960) shales and therefore did not recognize Harlton’s for- presented additional observations on the Mississippian mation names. As a result the Jackfork in Arkansas is age of the lower part of the Johns Valley and Jackfork divided into the Irons Fork Mountain Formation and in his study of the Lynn Mountain Syncline and greatly the overlying Brushy Knob Formation, and the contact influenced several students whose work on the Ouachi- between the two is the same as that between the Prai- tas was later published (Shelburne, 1960; Hart, 1963; rie Mountain and Wildhorse Mountains Formations of Seely, 1963; Fellows, 1964; Briggs, 1973). The age of Oklahoma. the Jackfork was clearly resolved by Gordon and Stone

13 OKLAHOMA GEOLOGICAL SURVEY

Game Refuge Sandstone (for- mation) Table II. — Informal Nomenclature The uppermost formation of of the Ouachita Mountains the Jackfork Group is the Game Refuge Sandstone, originally Atoka Formation named by Harlton (1959) for Fanshawe sandstone exposures near a state game ref- Red Oak sandstone uge on the southeast flank of the Panola sandstone Lynn Mountain Syncline. Harl- Diamond sandstone ton (1959) did not specify an Brazil sandstone exact location nor did he publish Bullard sandstone a measured section of the unit, Cecil sandstone rather, the location was chosen Shay sandstone because a name was available de- Foster sandstone spite better exposures elsewhere. Johns Valley Shale (formation) The only detailed map of the area Stapp conglomerate (1) (Pitt et al., 1982) does not show Jackfork Group the Game Refuge, rather, shows Hope sandstone only the upper (undivided) part Secor pay zone of the Jackfork Group there to Stanley Group be bounded by faults. Despite its Moyers Formation poor definition, the Game - Ref Chickasaw Creek tuff uge Sandstone has been used by Miller sand most subsequent workers in the Moyers siliceous shale; Siliceous shale member; Ouachita Mountains. Schoolhouse chert member (2) Tenmile Creek Formation Wesley Shale (formation) Middle Tenmile Creek siliceous shale; Middle siliceous shale bed; Battiest chert member (2); The type locality for the Wes- Smithville chert lentil ley Shale was precisely located Faith chert member (2) by Harlton (1938) on the north- Albion siliceous shale member; Tuskahoma siliceous east flank of an unnamed syncline shale (2); Middle Stanley siliceous shale; “near” (five miles southwest of) Albion Creek chert member (2); Campbell the village of Wesley. No mea- Creek siliceous shale(?) sured section of the unit in the Friendship chert member (2) type locality was published and Siliceous shale at base; Lower siliceous shale bed; Harlton’s (1938) description is Basal Ten Mile Creek Siliceous Shale; Black brief, except for those of several Knob Ridge chert member (2) petrographic thin sections. Harl- Mud Creek tuffs (upper and lower) ton (1938, p. 886) claimed the Beavers Bend tuff (3) Wesley was “one of the best de- veloped and most widespread of Blaylock Sandstone (formation) diagnostic shales … and is easily Beavers Bend illite (3) recognized and has been found at many localities in the Ouachita Mountain (sic) area,” but several Footnotes: (1) Stapp conglomerate member of Union Valley later workers were not able to Formation (Springer Group) recognized by OGS, not by distinguish it from the underlying USGS. (2) unit not recognized by USGS, recognized by Markham Mill Formation and/or OGS. (3) see text for discussion. overlying Game Refuge Sand-

14 FALL 2016

Figure 7. Wildhorse Mountain Formation, Jackfork Group. Location: along U.S. Highway 259 just west of Three Sticks Monu- ment. NE SW SE sec. 34, T. 2 N., R. 25 E. stone. Harlton (1938) also suggested the Wesley Shale was available” (p. 884). Harlton (1938) did not publish was present in the Choctaw Fault Zone and at the Pi- a measured section of the Markham Mill, and his de- netop schoolhouse and that its siliceous “character” ex- scription of the unit is brief. He separated the formation tended to the Arbuckle Mountains. No modern geologic from the underlying Prairie Mountain based on the pres- maps support this. Furthermore, he correlated the Wes- ence of a widespread shale that he named the Markham ley with the “Pennsylvanian Caney” of the Ardmore Mill siliceous shale. Harlton (1938, p. 884) suggested Basin and recommended replacing the latter with the that the Markham Mill Formation “is very widespread Wesley Shale. No work in the last 78 years has followed in distribution in the entire Ouachita Mountains and is this suggestion. easily recognized,” however, most subsequent workers were unable to distinguish it from the underlying Prairie Markham Mill Formation Mountain Formation and some from the overlying Wes- ley Shale. Harlton (1938) named the Markham Mill Forma- tion for a saw mill located along the northwest flank of Prairie Mountain Formation the Farris Syncline. He identified a second type local- ity along the western flank of the Round Prairie Syn- The Prairie Mountain Formation was named by cline that he appears to have preferred, but “no name Harlton (1938) presumably for a nearby Prairie Moun-

15 OKLAHOMA GEOLOGICAL SURVEY tain, but modern maps do not show a mountain with that Ouachita Mountains (Hendricks et al., 1947, “maroon name nearby. The type locality is closely located (north- shale member”; Cline, 1960; Shelburne, 1960; Hart, west flank of the Round Prairie Syncline) but the de- 1963; Fellows, 1964; Briggs, 1975; Pitt et al., 1982) scription of the unit is brief and is not accompanied by note the widespread nature of the Prairie Hollow Shale. a measured section. Harlton (1938) included the Prai- Seely (1963), however, stated that the unit was untrace- rie Hollow Shale Member, described below, within the able in the eastern frontal belt but suggested that it may Prairie Mountain Formation. Harlton (1938) also noted be covered by float. that a widespread marker bed that he called the Prairie Mountain siliceous shale was present at the base of the Wildhorse Mountain Formation formation. Many subsequent workers in the Ouachita Mountains were unable to identify this siliceous shale Harlton (1938) named the Wildhorse Mountain For- and some combined the Prairie Mountain with the upper mation (Figure 7) presumably for a nearby Wildhorse part of the Wildhorse Mountain on their maps. Mountain, but modern maps show his type section cross- ing Parker Mountain, Razorback Mountain, and White Prairie Hollow Shale Member Rock Mountain. The closest Wildhorse Mountain is about 20 mi to the east. Harlton’s (1938) type “locality” Harlton (1938) originally suggested that the Prairie is precisely located and described but about 70 percent Hollow Shale, named for Prairie Hollow on the west is covered, including the top and base. Harlton (1959, p. side of the Round Prairie Syncline, was the basal mem- 880) also identified a second, less precisely located type ber of the Prairie Mountain Formation and noted that it locality, but did not include a detailed description of it. was well exposed in Prairie Hollow (present on modern Despite these inadequacies, most modern studies of the maps and presumably the type locality), but he did not Jackfork Group recognize the Wildhorse Mountain For- specify a type locality nor present a measured section of mation as the basal unit in the group, but the upper part the unit. In 1959 he revised the position of the Prairie (above the Prairie Hollow Shale) is variably mapped as Hollow and placed it in the upper part of the Wildhorse undivided Wildhorse Mountain and Prairie Mountain. Mountain Formation. Most map-based studies in the

Harlton (1959) Cline (1960) Shelburne (1960) Briggs (1973) Hart (1963) Seely (1963) Pitt et al. (1982) Morris (1971)

Game Refuge Game Refuge Game Refuge Game Refuge Game Refuge Game Refuge Sandstone Sandstone Sandstone Sandstone Sandstone Sandstone Hendricks et al. (1947)

Siliceous Shale Wesley Shale Wesley Shale Wesley Shale Wesley Shale Wesley Shale Wesley Shale Member Brushy Knob Formation Upper Jackfork Markham Mill Markham Mill Markham Mill Markham Mill Markham Mill Markham Mill Group Formation Formation Formation Formation Formation Formation ? Prairie Prairie Prairie Prairie Prairie Prairie Mountain Mountain Mountain Mountain Mountain Mountain Formation Formation Formation Formation Formation Formation

? Upper Upper Upper Upper Wildhorse Wildhorse Wildhorse Wildhorse Wildhorse Mountain Mountain Mountain Mountain Mountain Formation Formation Formation Formation Formation

Maroon Shale Prairie Hollow Prairie Hollow Prairie Hollow Prairie Hollow Prairie Hollow Prairie Hollow Member Shale Member Shale Member Shale Member Shale Member Shale Member Wildhorse Shale Irons Fork Mountain Mountain Formation Formation Wildhorse Lower Lower Lower Middle & Lower Lower ? Mountain Wildhorse Wildhorse Wildhorse Wildhorse Jackfork Formation Mountain Mountain Mountain Mountain Group Formation Formation Formation Formation

Figure 8. Stratigraphic nomenclature of Jackfork Group as used by different authors. Thick black lines represent siliceous shale beds.

16 FALL 2016

Figure 9. Moyers Formation, Stanley Group. Location: east side of emergency spillway to Lake Carl Albert about two miles northwest of Talihina. NW NE NW sec. 2, T. 3 N., R. 21 E. Photograph taken by Brett Riley, University of Oklahoma.

Stanley Group (Chesterian to Osagean) (Hendricks et al., 1947; Cline, 1960; Pitt et al., 1982; Fellows, 1964; Hart, 1963; Seely, 1963; Briggs, 1973; Taff (1902) named the Stanley Group “Standley” Shelburne, 1960; Honess, 1923) and the top and base shale after the small town of Standley on the Kiamichi are generally well defined. Some early disagreements River in Pushmataha County. Purdue (1909) changed centered on the age of the unit, but most workers agreed the spelling to Stanley because the name of the village that the Stanley is Mississippian. Between 1934 and had changed and applied it to similar strata in Arkansas. 1959, however, Harlton (1934, 1938, 1959) greatly Only the upper part of the Stanley is exposed near the confused the entire issue of the age of the strata from village and Taff (1902) did not describe a type section the base of the Stanley Group to the base of the Ato- or type locality. Pitt et al. (1982), however, suggested ka Formation by suggesting that a “Bendian System” a type section for the Stanley Group extending from existed between the Mississippian and Pennsylvanian the south side of the Potato Hills south to the north and that a “Pushmataha Series” and/or a “Springer Se- flank of the Kiamichi Mountains. As noted by Pitt et ries” existed and included what is now the Johns Val- al. (1982, p. 19), however, “many units are covered ley, Jackfork, and Stanley. Harlton seems to have be- by colluvium and alluvium” (44 percent is covered or come so confused he misquoted himself (Suneson and mostly covered); in addition, two faults cross the line Hemish, 1994, p. 87). of section but are not noted in the detailed description. Harlton (1938) elevated the rank of the Stanley to The Stanley Group (Figure 9) is widely distrib- group and identified three formations – from bottom to uted throughout the Ouachita Mountains in Oklahoma top, the Tenmile Creek (shown as Ten Mile on his mea-

17 OKLAHOMA GEOLOGICAL SURVEY

Figure 10. Chickasaw Creek Shale, Stanley Group. Location: along Talimena Drive (Oklahoma Highway 1) east of U.S. High- way 259. NW SW NE sec. 1, T. 2 N., R. 25 E. sured section), Moyers, and Chickasaw Creek Shale. accurately summarized the basal Moyers siliceous shale Harlton (1938) divided the two older formations based as “laterally persistent over a wide area, (but) not pres- on the presence of a widespread siliceous shale that he ent everywhere ….”. Briggs (1973) also noted that the named and mapped as the Moyers siliceous shale (not sandstones and shales in the Tenmile Creek and Moy- member). The Chickasaw Creek consists of abundant ers Formations are similar and that the only difference siliceous shale beds and is lithologically distinct from between the two formations is that the Moyers contains the underlying Moyers and overlying Wildhorse Moun- more sandstone (35 to 40 percent) than the Tenmile tain Formation (Jackfork Group). Creek (15 percent). The Tenmile Creek and Moyers Formations have been accepted by most workers in the Ouachita Moun- Chickasaw Creek Shale (formation) tains, but in many places the basal Moyers siliceous shale is absent (or has not been recognized) and resul- Harlton (1938) named the Chickasaw Creek sili- tant maps show the two formations as undivided. Cline ceous shale (Figure 10) for exposures along Chickasaw (1960), Shelburne (1960), Fellows (1964), and Seely Creek on the north flank of an unnamed syncline about (1963) traced the shale eastward into Arkansas with four miles east of Stringtown. He precisely identified varying degrees of success; Shelburne (1960, p. 19) said the type locality but noted that the top and base of the the shale is “thin and discontinuous” and Fellows (1964, formation there is covered. Harlton (1938) identified a p. 29) said it is “fairly persistent.” Briggs (1973, p. 7) second type locality where the base was exposed but did

18 FALL 2016

Figure 11. Arkansas Novaculite. Location: small quarry at south end of Black Knob Ridge. SW SW SW sec. 13, T. 2 S., R. 11 E.

not note how much of the upper part was cov- ered. The Chickasaw Creek Shale is one of the most widespread and recognizable formations in the section in the Ouachita Moun- tains not only in Oklahoma but in Arkansas.

Moyers Formation

The Moyers Formation (Figure 9) is named for the village of Moyers, Pushmataha County, on the southeast nose of the Tuskahoma Syn- cline where Harlton (1938) precisely located a type section. The top and base of the formation are exposed and about 44 percent of the unit is not covered at the type section. As noted above, Figure 12. Blaylock Sandstone. Location: quarry used for construc- the name Moyers is used by most workers in the tion of emergency dam for . NE SW NW sec. 4, T. 5 Ouachita Mountains but is mapped as undivided S., R. 25 E. Photograph taken in 1988. with the Tenmile Creek Formation where the

19 OKLAHOMA GEOLOGICAL SURVEY basal Moyers siliceous shale is not present.

Tenmile Creek Formation

Harlton (1938) named the Ten- mile Creek Formation (shown as Ten Mile Creek in his figure 6) for exposures on the south nose of the Tuskahoma Syncline. Tenmile Creek is located in the valley on the west side of the syncline. Harl- ton (1938) precisely located a type section but the top of the unit there is not exposed and the base is over- lain by Cretaceous strata. About 63 percent of the type section is cov- ered. Harlton (1938) noted that sili- ceous shale beds are present in the Tenmile Creek and subsequent au- thors assigned them a variety of in- formal names; these are discussed below. Tuff beds are also present in the Tenmile Creek Formation; the most prominent one and the only one with a formal name is the Hat- ton Tuff Lentil, also described be- low.

BROKEN BOW UPLIFT

Stanley Group (Chesterian to Osagean)

A detachment zone (Arbenz, 2008) within the Stanley Group Figure 13. Bigfork Chert. Location: west end of Potato Hills along Cedar Creek. NE (probably within the Tenmile Creek NW NE sec. 31, T. 3 N., R. 20 E. Formation) separates the unit into an underlying complexly deformed Figure 14. sequence similar in complexity to Womble Shale. the Arkansas Novaculite and old- Location: Potato er rocks of the Broken Bow Up- Hills. Section- lift and an overlying sequence of Township- less-deformed rocks that conform Range to the broad synclines and sharp, unknown. typically thrust-faulted anticlines of the central belt of the Ouachita Mountains. The stratigraphic no- menclature of the Stanley Group in the lower sequence is the same as that in the upper.

20 FALL 2016

Figure 15. Collier Shale. Location: west shore of Broken Bow Lake. SE NW NE sec. 32, T. 4 S., R. 25 E. Photograph taken by Mike Willeby, Beavers Bend State Park.

Tenmile Creek Formation Arkansas Novaculite (formation) (Osagean to Lower Devonian) Hatton Tuff Lentil (note: lentil synonymous with lens as used by NACSN (2005)) Griswold (1892) first named the Arkansas Novacu- lite (also Arkansas stone) (Figure 11) as a kind of rock The Hatton Tuff Lentil was originally named by quarried for whetstones, but he also referred to the “no- Miser in an abstract in the Geological Society of Amer- vaculite formation.” Purdue (1909) mapped the Arkan- ica Bulletin in 1920 but more fully described by Miser sas Novaculite throughout a large part of the Arkansas and Purdue (1929). The unit was named for outcrops Ouachita Mountains and formally ranked it as a forma- along the Kansas City Southern railroad just south of tion, and Miser and Purdue (1929) suggested that the the town of Hatton in Polk County, Arkansas. Honess Caddo Gap area in Arkansas serve as the type locality. (1923, p. 150) believed only one tuff was present in Taff (1902) called the Arkansas Novaculite through the Stanley, in contrast to Miser (1921) who counted the Bigfork Chert at Black Knob Ridge the Talihina at least three and possibly as many as five tuff beds. chert (discussed below). Purdue (1909) recognized that Later workers confirmed Miser’s (1921) observations some of Taff’s Talihina chert there and similar strata in and named several, all of which are informal and are the Potato Hills were the same as those already named described below. in Arkansas (Hendricks et al., 1937), but Miser (1926) retained Taff’s nomenclature on the 1926 Geologic Map of Oklahoma. Honess (1923) followed Arkansas nomenclature and mapped the Arkansas Novaculite in

21 OKLAHOMA GEOLOGICAL SURVEY the Broken Bow Uplift area, and Miser (1929) did the 1902; Miser, 1926), but was replaced by Miser (1929) same in the Potato Hills as did Hendricks et al. (1937) and Hendricks et al. (1937). at Black Knob Ridge. In Arkansas the formation con- sists of three lithologic divisions (not members) (Mi- Womble Shale (formation) (Middle Ordovician) ser, 1918) but how these apply to the formation in Oklahoma is not agreed upon. The Womble Shale (Figure 14) was named by Miser (1918) for exposures near the town of Womble Missouri Mountain Shale (formation) (Upper Silu- (now Norman), Montgomery County, Arkansas and rian) was originally considered the upper part of Purdue’s (1909) Ouachita shale. The type section is at Crystal Purdue (1909) named the Missouri Mountain Mountain, Arkansas (USGS Geolex) but is unpub- Slate for exposures on Missouri Mountain (Missouri lished. Taff (1902) mapped the Womble as the String- Mountains on map) in southwestern Arkansas. He also town shale in the Black Knob Ridge area and Miser recognized that part of Taff’s (1902) Talihina chert at (1926, 1929) showed the Stringtown in the Potato Black Knob Ridge and in the Potato Hills was, in fact, Hills. Honess (1923) mapped the Womble throughout the Missouri Mountain. Like the Arkansas Novaculite, the Broken Bow Uplift area and first suggested (Hon- Honess (1923) applied the name Missouri Mountain in ess, 1930) that the shale beneath the Bigfork at Black the Broken Bow Uplift (slate), Miser (1929) in the Po- Knob Ridge and in the Potato Hills was the same age tato Hills (shale), and Hendricks et al. (1937) at Black as the Womble. Hendricks et al. (1937) followed Hon- Knob Ridge (shale). ess (1930) and replaced Stringtown with Womble at Black Knob Ridge. When the name Womble was first Blaylock Sandstone (formation) (Lower Silurian) applied to the oldest rocks in the Potato Hills is un- known. The Blaylock Sandstone (Figure 12) was named by Purdue (1909) for exposures at Blaylock Mountain Blakely Sandstone (formation) (Middle to Lower Or- in southwestern Arkansas. Honess (1923) mapped the dovician) same formation in the Broken Bow Uplift. The forma- tion is absent in the Potato Hills and at Black Knob Miser (1918) formally named the Blakely Sand- Ridge. stone for Blakely Mountain in Garland County, Arkan- sas, but did not identify or measure a type section. The Polk Creek Shale (formation) (Upper Ordovician) name was proposed by A.H. Purdue in a letter cited by Ulrich (1911) and Honess (1923) used the name in The Polk Creek Shale is named for Polk Creek his report on the Broken Bow Uplift. The Blakely and in Montgomery County, Arkansas (Purdue, 1909). older units are not exposed in the Black Knob Ridge or Like the Arkansas Novaculite and Missouri Mountain Potato Hills areas. Shale, it was included in the Talihina chert by Taff (1902) at Black Knob Ridge and in the Potato Hills on Mazarn Shale (formation) (Lower Ordovician) the 1926 Geologic Map of Oklahoma (Miser, 1926). Honess (1923) used Polk Creek Shale on his map of Miser (1918) named the Mazarn Shale for Maz- the Broken Bow Uplift, and Miser (1929) and Hen- arn Creek in Montgomery County, Arkansas; Purdue dricks et al. (1937) used Polk Creek in the Potato Hills (1909) included it in his Ouachita shale (lower part). and at Black Knob Ridge, respectively. In Oklahoma the unit is only exposed in the Broken Bow Uplift where Honess (1923) used the name and Bigfork Chert (formation) (Upper to Middle Ordovi- accepted the stratigraphy as identified in Arkansas. cian) Crystal Mountain Sandstone (formation) (Lower Or- Purdue (1909) named the Bigfork Chert (Figure dovician) 13) for its widespread exposure near the Big Fork (Big- fork on his map) Post Office in Montgomery County, The Crystal Mountain Sandstone was named for southwestern Arkansas. Like the younger formations the Crystal Mountains in Montgomery County, Arkan- through the Arkansas Novaculite, the Bigfork in Okla- sas by Purdue (1909). Honess (1923) accepted and fol- homa was originally called the Talihina chert (Taff, lowed the Arkansas nomenclature in his work in the

22 FALL 2016 Broken Bow Uplift area. INFORMAL STRATIGRAPHY OF THE OUACHITA MOUNTAINS Collier Shale (formation) (Lower Ordovician to Up- (INCLUDING THE ARKOMA BASIN) per Cambrian) INTRODUCTION The Collier Shale (Figure 15) was named for Col- lier Creek in Montgomery County, Arkansas by Pur- The currently used North American Stratigraphic due (1909). Like all the formation names in the Broken Code (NACSN, 2005) allows for the formal naming Bow Uplift, Honess (1923) followed the Arkansas no- of subsurface units, but the requirements for doing so menclature. are strict. For example, a description of the rock types that make up the unit are required; these descriptions SUMMARY OF FORMAL can come from examination of drill cuttings or cores. NOMENCLATURE NACSN (2005) only recommended, however, that the material used for the description be made available Most of the formal and most commonly used from a public repository. names in the Ouachita Mountains of Oklahoma might Because the Arkoma Basin is a major gas province be called “legacy” names that were established in the in Oklahoma and has been extensively drilled, a num- first decade of the twentieth century by Joseph Taff ber of reservoir units that are not exposed at the surface working in Oklahoma and Albert Purdue working in or have not been correlated with units exposed on the Arkansas before any “code of stratigraphic nomencla- surface have been named (Table II). The depositional ture” existed. As a result, type sections defined by the origins of these units and their reservoir characteristics original workers do not exist and, in some cases, type are related to Ouachita tectonism and they are there- localities or type areas are unclear. Later workers pro- fore discussed here. Because none of the subsurface posed type sections for previously named units, but in nomenclature is formalized, the following discussion most cases these should have been designated as refer- focuses on the origin of the name and is mostly based ence sections (NACSD, 2005). on information in a database available from IHS. The The nature and origin (and thus the name) of the two units that include informally named subsurface Johns Valley Shale and its relation to the Caney Shale petroleum reservoirs are the Atoka Formation and was controversial for many years but has been re- Jackfork Group; the named reservoirs are described solved, although the extent of the Johns Valley in the from youngest to oldest. northern part of the Ouachita Mountains and the pos- In addition to the subsurface units in the frontal sible existence of large, coherent slide masses that ap- belt and Arkoma Basin, a number of surface units have pear to be stratigraphically autochthonous remains a been given names (Table II) and in some cases a type topic worthy of further research. section and/or type locality, but the regional extent Bruce Harlton’s names for the Jackfork Group for- of the unit is questionable. Some of these informally mations confused stratigraphers for decades. Harlton named units are referred to in only one paper and oth- did not follow the newly published code but, more im- ers are referred to by several authors but are given dif- portantly, he mistakenly suggested that key siliceous ferent names. marker beds that he used as formation boundaries were widespread. Tom Hendricks wisely rejected the sug- FRONTAL BELT gestion in his mapping and, much later, so did William Pitt and his colleagues. In contrast, Louis Cline and his Atoka Formation students attempted to use Harlton’s names but, lack- ing the siliceous shales, were forced to combine units. Fanshawe sandstone Had they not followed Harlton’s suggestion, they may The Frankfort No. 1 Hulsey (SE sec. 18, T. 6 N., R. have divided the Jackfork into four mappable units 22 E.) was the first well to identify the Fanshawe sand- that most of them recognized – from top to bottom, the stone in the subsurface. The well spudded on August Game Refuge Sandstone, a “middle Jackfork forma- 8, 1960, drilled to 12,465 ft total depth (TD), and was tion,” Prairie Hollow Shale, and Wildhorse Mountain completed as a Red Oak gas well on August 1, 1961. Formation. The well tested the Cromwell, Spiro, Red Oak, and Fanshawe sandstones; all showed gas. Drilling reports show the top of the Fanshawe at 6296 ft drilled depth

23 OKLAHOMA GEOLOGICAL SURVEY and a tested interval from 6296 ft to 6320 ft. The unit was probably named after the town of Fanshawe, which is located about seven miles south-southeast of the well. The Fanshawe currently produces from a number of gas fields throughout the Arkoma Basin, including Bokoshe South, Red Oak – Norris, Wilburton, and Haileyville Southwest (field locations after Pritchett, 2015).

Red Oak sandstone

The first well to produce gas from a unit identified as the Red Oak sandstone is the Mid- west No. 1 Kane (NW sec. 8, T. 6 N., R. 23 E.). The well spudded on April 23, 1955, and was Figure 16. Miller tar sand. Location: Redden oil field. SE NW NE sec. completed on July 12, 1966 in the Red Oak. 9, T. 1 S., R. 14 E. Well records show the top of the Red Oak at May 12, 1983, TD’d at 12,890 ft., and was completed 8600 ft and a perforated interval from 8606 ft to 8705 as a gas well in the middle Atoka on January 16, 1984. ft. The well probably was named for the town of Red The operator identified the top of the Diamond sand- Oak, which is about 11 mi west-southwest of the well. stone at 6680 ft drilled depth. The origin of the name The Red Oak sandstone is one of the most prolific of the sandstone is unknown. There are no geographic Atoka Formation sandstones in the Arkoma Basin and features in the area with the name Diamond. produces gas in the Wilburton, Kinta, Red Oak – Nor- The Diamond sandstone produces gas from about ris, and Bokoshe South fields, as well as in wells in ten wells in the Panola field. other fields in the southern part of the basin.

Brazil sandstone Panola sandstone The Pan American No. 1 C.L. Gould was the first The Sunset No. 1 Fisherman (NE sec. 34, T. 6 N., well to identify the Brazil sandstone (SE sec. 35, T. 8 R. 20 E.) was the first well to identify the Panola sand- N., R. 23 E.). It spudded on September 22, 1966 and stone in the subsurface. The well spudded on June 12, drilled to a TD of 12,360 ft. It was completed as a gas 1969, drilled to 13,200 ft TD, and was completed on well in the Red Oak sandstone on February 16, 1967. September 30, 1969. The well tested minor gas in the In addition to the Red Oak, the well tested some gas in Red Oak sandstone. The top of the Panola sandstone the Brazil sandstone from 8140 ft to 8160 ft. Drilling was reported at 9304 ft drilled depth. The sandstone reports noted the top of the Brazil at 8018 ft drilled was named for the town of Panola, located about two depth. The sandstone is probably named for the town miles southwest of the well. Gas was discovered in the of Brazil, which is located about four miles east of the unit about 15 years later immediately southwest of the well. town. The Brazil sandstone is productive in the Bokoshe The Panola currently produces from the Red Oak South, Red Oak – Norris, and Haileyville Southwest – Norris, Panola, and McAlester Southeast gas fields, fields and in scattered wells in other fields in the south- as well as scattered locations throughout the southern ern part of the Arkoma Basin. part of the Arkoma Basin. Bullard sandstone Diamond sandstone The Tenneco No. 1-13 Heitner (NE sec. 13, T. 5 The first well to identify the Diamond sandstone in N., R. 19 E.), was the first well to identify the Bullard the subsurface is the Williford No. 1-7 Butzer, located sandstone. The well spudded on February 14, 1984, in SE sec. 7, T. 5 N., R. 20 E. The well spudded on drilled to 13,000 ft TD, and was completed as a gas

24 FALL 2016 well in the Diamond, Panola, and Bullard sandstones. and the Foster is reported to produce gas only from this Drilling reports show Bullard perforations from 9964 well. ft to 10,011 ft and that the top of the Bullard was at The Foster sandstone consists of early Atokan flu- 9962 ft drilled depth. The origin of the name Bullard is vial sandstones eroded into late Morrowan strata along unknown; there are no geographic features in the area the northern margin of the Arkoma Basin (Lumsden et with the name Bullard. al.,1971); the channels transported sand from north to The Bullard sandstone produces gas from about 12 south where the sand was reworked by marine process- wells in the Panola gas field. es. These reworked sandstones are partly coeval with and partly overlie the Foster and are recognized as the Cecil sandstone Spiro Sandstone (described above). Thus, the Foster sandstone represents, in part, the landward equivalent The first report of the Cecil sandstone in Okla- of the Spiro. Only six other wells identified the Foster homa is from the Diamond Shamrock No. 1-5 Lear- sandstone in their drilling reports and none of them re- Vliet Trust well (NW sec. 5,T. 9 N., R. 27 E.). The well ported the Foster to be reservoir strata. spudded on October 11, 1975, drilled to 6891 ft TD, and was completed as a Cecil gas producer. The top of Johns Valley Shale the Cecil was reported at 5854 drilled depth, and Cecil perfs were from 5926 ft to 5938 ft. Stapp conglomerate The Cecil sand is a productive sandstone in the Ce- cil gas field named for the town of Cecil, Arkansas. In Harlton (1938) first named the Stapp conglom- Oklahoma the Cecil produces mostly from the Panola erate for exposures along the Kansas City Southern and Poteau Southeast fields. Railway about one mile south of the village of Stapp and identified it as a member of the Morrowan Union Shay sandstone Valley Formation. All later workers consider the Stapp conglomerate to be a facies of the Johns Valley Shale The Unit No. 1 Cox (NE sec. 8., T. 5 N., R. 20 E.) (summarized in Suneson and Hemish, 1994b; Suneson first identified the Shay sandstone. The well spudded et al., 2005, p. 59). on February 11, 1983, drilled to 12,400 ft TD and was completed in the Shay on July 27, 1983 as a gas well. Jackfork Group The Shay was reported at 12,120 ft to 12,202 ft drilled depth and all perforations were within that interval. Hope sandstone The origin of the name of the unit is unknown; there are no geographic features in the area named Shay. Cunningham and Namson (1994) proposed the in- The Shay sandstone produces gas from about 12 formal name Hope sandstone for a reservoir unit in the wells in the Panola field as well as a single well in the lower part of the Jackfork Group in two wells in the Red Oak South field. Buffalo Mountain and Talihina Northwest gas fields. Very little has been published on the subsurface stra- Spiro Sandstone tigraphy of this area and the name has not been used by other geologists. The Spiro Sandstone is considered a formal unit in this paper and is discussed above. Secor pay zone

Foster sandstone Montgomery (1996) used the name Secor pay zone to describe a reservoir interval in three wells the Jack- The Humble 1 Summings Estate (SW sec. 9, T. 8 fork in the Buffalo Mountain and Talihina Northwest N., R. 22 E.) is the first well to identify the Foster sand- fields. Like the Hope sandstone above, this name is not stone. It spudded on December 20, 1964, drilled to used. 6250 ft TD, and was completed in the Foster on March 4, 1965. The top of the Foster was drilled at 6025 ft, was cored from 6108 ft to 6201 ft, and perforated from 6105 ft to 6191 ft. The origin of the unit is unknown

25 OKLAHOMA GEOLOGICAL SURVEY

Stanley Group Harlton (1938) on the east side of the school in his type section of the Moyers Formation. Moyers Formation Tenmile Creek Formation Chickasaw Creek tuff Middle Tenmile Creek siliceous shale, Middle sili- The Chickasaw Creek tuff is the youngest and ceous shale bed, Battiest chert member, Smithville least-studied pyroclastic deposit in the Stanley Group. chert lentil It appears to have been first identified in a Ph.D. disser- tation (Laudon, 1959). Hart (1963) and Seely (1963) Harlton (1938) identified several siliceous shale identified a tuff or tuffaceous sandstone in the Chicka- beds in the Tenmile Creek Formation that he consid- saw Creek Shale on the south flank of the Simmons ered to be widespread; later workers had varying de- (Winding Stair) Mountain Syncline and north flank of grees of success identifying the same beds elsewhere in the Eagle Gap Syncline, respectively, but they did not the Ouachita Mountains. The youngest siliceous shale name it. Mose (1969) first used the name Chickasaw is labeled Middle Ten Mile Creek siliceous shale on Creek Tuff (capital “T”); later workers (e.g., Niem, his map (figure 6), but Harlton (1938) noted that it is 1977; Loomis et al., 1994; Shaulis et al., 2012) all used “poorly represented in known sections” (p. 868). Cline a lower-case “t,” although it is not clear whether or not (1960) named the same unit the Middle siliceous shale the use of a capital vs. lower-case “t” was intention- bed and noted that it is a widespread marker bed in the al. Morris (1974) stated that the “tuffaceous interval” Tenmile Creek Formation throughout the Ouachitas, within the Chickasaw Creek Shale was present for 100 but he did not map it on his map of the Lynn Mountain mi in the northern part of the Ouachitas in Oklahoma Syncline. Shelburne (1960) named the same unit the and Arkansas. Battiest chert member and noted many localities in the Boltukola Syncline area where it is well exposed. He Miller sand described it well but did not propose a type section. Pitt et al. (1982) also named the unit the Battiest chert The Miller sand (Figure 16) is a general name giv- member and noted that it is the most “persistent” of en to shallow oil-productive sandstones in the Stanley the Tenmile Creek siliceous shales. Miser and Honess Group in the Redden oil field (Campbell, 1990). The (1927) named the unit the Smithville chert lentil, how- IHS database lists four wells spudded between 1930 ever, that name had been preempted (Branson, 1957) and 1943 that produced oil; all were located on prop- and was never used in subsequent work. erty owned by E.P. Miller, after whom the sandstones were named. Faith chert member

Moyers siliceous shale, Siliceous shale member, Pitt et al. (1982) identified and named the Faith Schoolhouse chert member chert member in the Tenmile Creek Formation. They published a type locality but did not provide a mea- Harlton (1938) identified and described a siliceous sured section; in addition, the type locality was outside shale that he used to mark the base of the Moyers For- the area of their report so there is no record of how mation. He located the shale within his type section of continuous the unit is. the Moyers Formation, but he did not describe the un- derlying strata and the strata over the shale were cov- Albion siliceous shale member, Tuskahoma siliceous ered. He identified a “second type locality” only gener- shale, Middle Stanley siliceous shale, Albion Creek ally located (probably along the road in the identified chert member, Campbell Creek siliceous shale(?) section) but did not describe it. Cline (1960) mapped the same siliceous shale bed at the base of the Moyers Harlton (1947) briefly mentioned using the name and called it a member of the Moyers Formation; he Albion siliceous shale member for a unit he observed also noted that the siliceous shale member was “wide- in the lower part of the Tenmile Creek Formation on spread” (p. 37). Pitt et al. (1982) renamed the unit the the east side of Black Knob Ridge. Goldstein and Hen- Schoolhouse chert member for outcrops just west of dricks (1953) renamed the unit the Tuskahoma sili- the schoolhouse in Moyers, but did not describe the ceous shale for its type section near the town of Tuska- section. These outcrops are close to those used by homa; they noted that the section had been measured

26 FALL 2016 by T.A. Hendricks and P. Averitt, who called it the tion by Hill and published a very generalized measured Middle Stanley siliceous shale (Tulsa Geological So- section of the tuff at its type section. The tuff is named ciety, 1947, p. 35). Pitt et al. (1982) renamed the unit for the state park in which the type section occurs, and the Albion Creek chert member and identified a type the park is named for an abrupt bend in the Mountain area near Albion Creek on the east end of the Potato Fork River next to land originally owned by John T. Hills; however, they did not measure a type section of Beavers, an early settler in the area. Although a num- the unit, nor did they describe it. ber of workers have studied the unit, it is considered Hendricks and Averitt (Tulsa Geological Society, informal. Geolex notes that the name is “… considered 1947, p. 35) noted that “this siliceous shale may be the invalid,” that it has been abandoned or is no longer same as the Campbell Creek siliceous shale of Harl- used, and that it is “uncertain if used by the Oklahoma ton,” but there is no mention of a Campbell Creek in Geological Survey. Not used by the USGS”. any of Harlton’s published work. Blaylock Sandstone Friendship chert member Beavers Bend illite Pitt et al. (1982) identified a chert bed they named the Friendship chert member of the Tenmile Creek Mankin and Dodd (1963) named the Beavers Bend Formation. The chert is named after a school about illite for an outcrop in the Blaylock Formation in Bea- four miles west of the type section, which they pre- vers Bend State Park near Broken Bow. They located cisely located. Pitt et al. (1982, p. 75) described the the outcrop precisely and published a brief description chert at its type section, but the description is cursory but did not attempt to map the unit beyond where they and 45 percent of the section is covered. identified it. Geolex precedes its listing with a [?] and notes that the unit is “not synopsized to date.” Siliceous shale at base, Lower siliceous shale bed, Black Knob Ridge chert member SUMMARY OF INFORMAL NOMENCLATURE Like all the other siliceous shale and/or chert units in the Stanley Group, the lowest siliceous unit has been Natural gas reservoir units occur in the Atoka For- given a number of names. Harlton (1938) informally mation and Jackfork Group; except for the prolific called the unit the siliceous shale at the base. Gold- Spiro Sandstone that is easily recognized on the sur- stein and Hendricks (1953) called it the Basal Ten Mile face and in the subsurface, their names should remain Creek siliceous shale, Cline (1960) called it the Lower informal as the current code recommends. siliceous shale bed, and Pitt et al. (1982) named it the Bruce Harlton recognized a number of siliceous Black Knob Ridge chert member. Despite its suppos- shale beds in the Tenmile Creek Formation, as did edly formal name, Pitt et al. (1982) did not locate or other later workers, but a failure to correlate them describe a type section. throughout the Ouachita Mountains has made naming them impossible. The reasons for this failure are le- Mud Creek tuffs (upper and lower) gitimate: the Stanley is structurally complex, the ex- posures are poor, and it is possible that the shales are Niem (1977) named the upper and lower Mud not continuous. These siliceous shales should be reex- Creek tuffs for exposures near the town of Bethel. amined; perhaps the trace-element geochemistry of the (Mud Creek is about six miles south of Bethel.) He did shales could be used to correlate them. not publish measured sections of the tuffs and subse- The Hatton Tuff Lentil is an important unit be- quent workers (e.g., Loomis et al., 1994; Shaulis et al., cause it provides some data on the plate tectonic his- 2012) differ on whether to consider it an informal or tory of what some researchers call the Ouachita Basin. formal unit within the Tenmile Creek Formation. The The Beavers Bend, two Mud Creek, and Chickasaw unit is not shown in the USGS Geolex database. Creek tuffs deserve more study and, assuming those studies can meet the recommendations of the current Beavers Bend tuff code, recognition as formal units.

Niem (1977) stated that the Beavers Bend tuff was first identified and named in 1967 in a Ph.D. disserta-

27 OKLAHOMA GEOLOGICAL SURVEY OBSOLETE NAMES

INTRODUCTION Table III. — Obsolete Nomenclature

Two names established by Taff (1902) were Brushy Creek chert later determined to consist of several formations Talihina chert that were identified in Arkansas. Units named Stringtown shale early in the geological investigations of an area are typically subdivided into others upon more detailed mapping and stratigraphic study and the Ouachita Mountains are no exception. These ear- ly names are now considered obsolete (Table III) Womble Shale in 1938 (Wilmarth, 1938), but when and are shown in lower-case letters below. Womble was first applied to the oldest rocks in the Po- tato Hills is unknown. Brushy Creek chert INCORRECT NOMENCLATURE Ulrich (1927, figure 2) named the Brushy Creek chert for limestone and chert outcrops underlying the INTRODUCTION Woodford Shale along Brushy Creek. Miser (1934, footnote, table 1) replaced the name with Pinetop For the purposes of this paper, incorrect names Chert after the nearby Pinetop School because the (Table IV) are those that were proposed by geologists name Brushy Creek was preempted. who flagrantly ignored the existing code of stratigraph- ic nomenclature or miscorrelated units from other tec- Talihina chert tonic provinces. In this paper these names are shown in lower-case letters. (Names shown in Table IV with Taff (1902) originally used Talihina chert to refer capital letters are accepted in other geologic provinces to a sequence of cherts and shales in the Black Knob but should not be used in the Ouachitas.) In the case of Ridge area of Atoka County. He presumably named the stratigraphy of the Oklahoma Ouachita Mountains, the sequence after similar rocks he had observed in the nomenclature proposed by two geologists – Bruce the Potato Hills, just west of the town of Talihina in Harlton and William Pitt – must be viewed with a great Le Flore County. Purdue (1909) mapped and subdi- deal of skepticism. vided the sequence into the Bigfork Chert, Polk Creek Shale, Blaylock Sandstone, Missouri Mountain Slate, Lynn Mountain formation and Arkansas Novaculite in Arkansas. Honess (1923) followed Arkansas nomenclature in the Broken Bow The Lynn Mountain formation was used by Pitt et Uplift area, Miser (1929) replaced Talihina chert with al. (1982) to include the same strata mapped by others the Arkansas nomenclature in the Potato Hills, and in Pushmataha County (e.g., Cline, 1960) as the Atoka Hendricks et al. (1937) did the same in the Black Knob Formation but gave no reason why they did so. March- Ridge area. er and Bergman (1983) mapped all strata younger than the Johns Valley Shale and the Wapanucka Limestone Stringtown shale south of the Choctaw Fault as Lynn Mountain and lim- ited using Atoka Formation to those rocks north of the Taff (1902) identified the Stringtown shale beneath Choctaw Fault older than the Hartshorne Formation. the Talihina chert in the Black Knob Ridge area and Suneson (1987) reviewed the objections to using the named it for the town of Stringtown. Miser (1918) name Lynn Mountain formation and recommended mapped the same formation in Arkansas and named abandoning it; a key argument was that the age of the it the Womble Shale. Honess (1923) accepted Miser’s Lynn Mountain was poorly constrained and based on nomenclature and applied it to his work in the Broken “tentative” identification of Morrowan palynomorphs Bow Uplift area. Ulrich (1927) equated the Stringtown outside of Pushmataha County and “Atoka palyno- and Womble Shales, and Hendricks et al. (1937) used morphs of Desmoinesian age” (Pitt et al., 1982, p. 35) the name Womble in their work on Black Knob Ridge. in Pushmataha County. The detailed palynological The USGS formally replaced Stringtown shale with

28 FALL 2016 Limestone and above the Primrose For- mation in the frontal belt of the Ouachitas and correlated it with a similar shale in the Ardmore Basin. Marcher and Berg- Table IV. — Incorrect Nomenclature man (1983) also showed the Limestone Gap in the northwestern part of the moun- Lynn Mountain formation (1) tains. Hendricks et al. (1947) mapped the Barnett Hill formation (2) same unit as the Wapanucka Limestone. Limestone Gap shale (1) Few subsequent workers followed Harl- Primrose formation (2) ton’s (1938) use of the term and most Round Prairie formation (1) include Harlton’s Limestone Gap in the Union Valley Sandstone Wapanucka Limestone. Goddard Shale (2) Delaware Creek Shale (2) Primrose formation Lukfatah sandstone (1) Harlton (1938) correlated the well- studied Primrose Formation in the Ar- Footnotes: no number – unit recognized by USGS dmore Basin with what Taff (1901) and OGS. (1) unit not recognized by USGS, recog- called the Chickachoc chert lentil. Harl- nized by OGS. (2) unit recognized by USGS and ton (1938, p. 901) noted a good expo- OGS in Arbuckle Mountains, not in Ouachita Moun- sure south of Blanco that Hendricks et tains. al. (1947) mapped as Wapanucka Lime- stone. Like the Limestone Gap, recent workers include the Primrose with the Wapanucka. work was never published. The name Atoka is so well recognized in the geo- Round Prairie formation logical literature on the Ouachita Mountains and Ar- koma Basin that it should be retained and Pitt et al.’s The origin of the name Round Prairie is related to (1982) and Marcher and Bergman’s (1983) changes the origin of the Caney olistoliths in the Johns Valley should be ignored. Shale and Harlton’s (1938) use of the Bendian period between the Mississippian and Pennsylvanian. Harlton Barnett Hill formation (1938) divided what others recognized as the Johns Val- ley Shale in Johns Valley into three formations based Harlton (1938) first proposed using the name Bar- on the age of the shale containing the olistoliths; from nett Hill for Morrowan limestones and cherts underly- bottom to top, the Caney, Wesley, and Round Prairie. ing Desmoinesian (author’s italics) Atoka Formation. Harlton’s (1938) Caney probably is a large olistolith, He correlated it with the Barnett Hill Formation near possibly a slide mass. The Wesley probably consists Clarita on the eastern end of the Arbuckle Mountains mostly of Morrowan shale, and the Round Prairie is and separated it from the underlying Wapanucka Lime- more typical Johns Valley and consists of numerous stone based on fossils and the character of the chert boulder beds. Harlton (1938) did not, however, extend nodules. Laudon (1958) noted that the Barnett Hill is his proposed stratigraphy into his type locality of the lithologically similar to the Atoka Formation and, as a Round Prairie where it is shown overlying the Union result, Harlton’s (1938) suggestion was never accepted Valley sandstone. Hendricks et al. (1947) mapped by later workers. Harlton’s (1938) Round Prairie type locality as Johns Valley Shale and no recent workers in the Ouachitas Limestone Gap shale use Round Prairie.

Harlton (1938, figure 1) proposed using the name Union Valley sandstone Limestone Gap for a shale below the Wapanucka Harlton (1938) correlated the Union Valley sand-

29 OKLAHOMA GEOLOGICAL SURVEY stone in the Arbuckle Mountains with a unit of the SUMMARY OF INCORRECT same age in the Ouachita Mountains. Later workers, NOMENCLATURE as well as Harlton (1959), recognized this unit as the uppermost member of the Jackfork Group – the Game William Pitt and his colleagues ignored the code Refuge Sandstone. and named and renamed many of the units in Push- mataha County. Renaming one of the most widespread Goddard shale units in the area – the Atoka Formation – without expla- nation was an egregious error that was repeated on the Marcher and Bergman (1983) mapped the Goddard 1:250,000 geologic map of the McAlester-Texarkana shale above the Delaware Creek shale and below the quadrangles (Marcher and Bergman, 1983), a map co- Limestone Gap shale (Wapanucka Limestone of Hen- produced with the USGS. That same map attempted to dricks et al., 1947) in the frontal belt of the Ouachita correlate and rename shales in the Ouachita Mountains Mountains. The type locality of the Goddard is in the with well-studied shales in the Ardmore Basin without Arbuckle Mountains and it has been widely studied in evidence. Tom Hendricks’ 1947 USGS map retain the the Ardmore Basin. Hendricks et al. (1947) mapped correct nomenclature of the frontal belt. the Goddard in the frontal belt as Springer Formation, In addition to his mistaken suggestion regarding and this has been accepted by most recent workers. the extent of the siliceous shales in the Jackfork Group that he saw at the southern end of the Tuskahoma Syn- Delaware Creek shale cline, Bruce Harlton not only attempted to add another geologic period between the Mississippian and Penn- Marcher and Bergman (1983) mapped the Dela- sylvanian, he forced well-studied units in the Ardmore ware Creek Shale near or at the base of numerous Basin into his timescale and then applied those Ard- thrust sheets in the frontal belt of the Ouachita Moun- more Basin names to what he thought were correlative tains. The Delaware Creek Shale is widely recognized units in the Ouachitas. He revised and re-revised his in the Arbuckle Mountains as a member of the Caney nomenclature more than once. Fortunately, no Ouachi- Shale. Hendricks et al. (1947) mapped the same unit ta Mountains workers accepted his stratigraphy. in the Ouachitas as the Caney Shale, where it is not William Pitt’s attempt to find the oldest formation divided into individual members. Some of the more in the Ouachita perhaps led him to overlook the sedi- extensive Caney (Delaware Creek) outcrops probably mentary structures that would have been evidence that are stratigraphically autochthonous units at the base of even subhorizontal strata can be overturned. thrust sheets; others probably are slide masses within the Johns Valley Shale. RECOMMENDATIONS

Lukfata sandstone 1. Elevate the Spiro Sandstone of the Atoka Formation to formal member status; Pitt (1955) identified what he thought was the old- est formation in the Ouachita Mountains and named it 2. Reexamine the stratigraphy and nomenclature of the the Lukfata sandstone. He precisely located a type sec- Jackfork Group in light of regional studies; tion along Lukfata Creek and described it, although the (physically) basal contact was not exposed. Pitt (1955) 3. Determine the extent of the tuffs in the Stanley divided the formation into three members based on dif- Group, establish principal reference sections, and el- ferences in lithology, and the concept of a pre-Collier evate to formal member status; Shale formation in the Ouachitas was widely cited in the literature until Goldstein (1975) questioned it. 4. Determine the extent and continuity of siliceous Shortly afterwards, Visher et al. (1978, p. 130-131) shales in the Stanley Group using modern analytical noted that sedimentary structures within the sandstone techniques and identify with a single informal or for- indicated it overturned, and conodonts studied by mal name. Repetski and Ethington (1977) showed that the Luk- fata was younger, rather than older, than the Collier ACKNOWLEDGEMENTS Shale. First and foremost, I would like to thank Brittany

30 FALL 2016 Pritchett, chair of the OGS Stratigraphic Nomencla- available: Oklahoma Geology Notes, v. 17, p. 99-108. ture Committee, for organizing a group of geologists willing to take on a seemingly endless task of address- Briggs, G., 1973, Geology of the eastern part of the ing Oklahoma’s stratigraphic nomenclature problems. Lynn Mountain Syncline, Le Flore County, Oklahoma: In addition, she reviewed an early version of this re- Oklahoma Geological Survey Circular 75, 34p. port and made many excellent suggestions; the paper is very much better as a result of her review. Brian Car- Campbell, J.A., 1990, Stop 10, Redden oil field, in dott read a later version of the paper and caught sev- N.H. Suneson, J.A. Campbell, and M.J. Tilford, eds., eral important changes that had to be corrected. Good Geology and resources of the frontal belt of the west- eyes, Brian. I would also like to thank OGS employees ern Ouachita Mountains, Oklahoma: Oklahoma Geo- Jim Anderson for his cartographic work on Figures 1 logical Survey Special Publication 90-1, p. 62-65. and 2 and Ted Satterfield for his editorial and layout work. Because my work in the Ouachitas ended years Cline, L.M., 1956, Some stratigraphic studies of the ago before the age of digital photography, I used many Mississippian and Pennsylvanian rocks of the Ouachi- photographs taken by others. They are given credit in ta Mountains, Oklahoma: Tulsa Geological Society the captions and I would like to thank them for their Digest, v. 24, p. 100-106. permission to use their work. Finally, a shout-out to OGS Director Jerry Boak for resurrecting Oklahoma Cline, L.M., 1960, Stratigraphy of the late Paleozoic Geology Notes. rocks of the Ouachita Mountains, Oklahoma: Oklaho- ma Geological Survey Bulletin 85, 113p. REFERENCES CITED Cunningham, R.D., and J. Namson, 1994, Map-scale ACSN, 1961, Code of stratigraphic nomenclature: fold in the Ti Valley thrust sheet of the Ouachita fold- Bulletin of the American Association of Petroleum and-thrust belt, Latimer and Le Flore Counties, Okla- Geologists, v. 45, p. 645-665. homa, in N.H. Suneson and L.A. Hemish, eds., Geol- ogy and resources of the eastern Ouachita Mountains ACSN, 1970, Code of stratigraphic nomenclature (2nd frontal belt and southeastern Arkoma Basin, Oklaho- ed.): American Association of Petroleum Geologists, ma: Oklahoma Geological Survey Guidebook 29, p. 45p. 243-248.

Amsden, T.W., 1983, Coelospira concava (Hall) from Elias, M.K., 1956, Upper Mississippian and Lower the Pinetop Chert (Early Devonian), Ouachita Moun- Pennsylvanian formations of south-central Oklahoma, tains, Oklahoma: Journal of Paleontology, v. 57, p. in I.C. Hicks, J. Westheimer, C.W. Tomlinson, D.M. 1244-1260. Putnam, and E.L. Selk, eds., Petroleum geology of southern Oklahoma: A symposium sponsored by the Arbenz, J.K., 2008, Structural framework on the Ardmore Geological Society, volume 1: American As- Ouachita Mountains, in N.H. Suneson, ed., Strati- sociation of Petroleum Geologists, Tulsa, Oklahoma, graphic and structural evolution of the Ouachita p. 56-134. Mountains and Arkoma Basin, southeastern Oklahoma and west-central Arkansas: Applications to petroleum Elias, M.K., and C.C. Branson, 1959, Type section of exploration: 2004 field symposium: Oklahoma - Geo the Caney Shale: Oklahoma Geological Survey Circu- logical Survey Circular 112A, p. 4-40. lar 52, 24p.

Ashley, G.H., M.G. Cheney, J.J. Galloway, C.N. Fellows, L.D., 1964, Geology of the western part of Gould, C.J. Hares, B.F. Howell, A.I. Levorsen, H.D. Winding Stair Range, Latimer and Le Flore Counties, Miser, R.C. Moore, J.B. Reeside, Jr., W.W. Rubey, Oklahoma: Oklahoma Geological Survey Circular 65, T.W. Stanton, G.W. Stose, and J.B. Twenhofel, 1933, 102p. Classification and nomenclature of rock units: Bulletin of the American Association of Petroleum Geologists, Goldstein, A., Jr., 1975, Geologic interpretation of Vi- v. 17, p. 843-863. ersen and Cochran’s 25-1 Weyerhaeuser well, McCur- tain County, Oklahoma: Oklahoma Geology Notes, v. Branson, C.C., 1957, Old stratigraphic names made 35, p. 167-181.

31 OKLAHOMA GEOLOGICAL SURVEY

Goldstein, A., Jr., and T.A. Hendricks, 1953, Siliceous Harlton, B.H., 1938, Stratigraphy of the Bendian of the sediments of Ouachita facies in Oklahoma: Bulletin of Oklahoma salient of the Ouachita Mountains: Bulletin the Geological Society of America, v. 64, p. 421-442. of the American Association of Petroleum Geologists, v. 22, p. 852-914. Goldston, W.L., Jr., 1922, Differentiation and structure of the Glenn Formation: Bulletin of the American As- Harlton, B.H.,1947, The siliceous shale members of the sociation of Petroleum Geologists, v. 6, p. 5-23. Stanley – Jackfork deposits in the Ouachita Mountains, Oklahoma, with notes on the “Johns Valley” shale, in Gordon, M., Jr., and C.G. Stone, 1977, Correlation of J.C. Maher, ed., Field conference in western part of the the Carboniferous rocks of the Ouachita trough with Ouachita Mountains in Oklahoma: Tulsa Geological those of the adjacent foreland, in C.G. Stone, ed., Sym- Society guidebook, p. 41-46. posium on the geology of the Ouachita Mountains: Ar- kansas Geological Commission Miscellaneous Publica- Harlton, B.H., 1959, Age classification of the upper tion 13, p. 70-91. Pushmataha series in the Ouachita Mountains, in L.M. Cline, W.J. Hilsewick, and D.E. Feray, eds., The geol- Grayson, R.C., Jr., 1979, Stop descriptions – fifth day,in ogy of the Ouachita Mountains, a symposium: Dallas P.K. Sutherland and W.L. Manger, eds., Mississippian – Geological Society and Ardmore Geological Society, p. Pennsylvanian shelf – to – basin transition, Ozark and 130-139. Ouachita regions, Oklahoma and Arkansas: Oklahoma Geological Survey Guidebook 19, p. 67-76. Hart, O.D., 1963, Geology of the eastern part of Wind- ing Stair Range, Le Flore County, Oklahoma: Oklaho- Grayson, R.C., Jr., and L.W. Hinde, 1993, Lower Atoka ma Geological Survey Bulletin 103, 87p. Group (Spiro sand) stratigraphic relationships, depo- sitional environments, and sand distribution, frontal Hemish, L.A., N.H. Suneson, and C.A. Ferguson, 1990, Ouachita Mountains, Oklahoma, in K.S. Johnson and Geologic map of the Red Oak 7.5’ quadrangle, Latimer J.A. Campbell, eds., Petroleum-reservoir geology in the County, Oklahoma: Oklahoma Geological Survey Geo- southern Midcontinent, 1991 symposium: Oklahoma logic Map OGQ-6, scale 1:24,000. Geological Survey Circular 95, p. 216-244. Hendricks, T.A., 1939, Geology and fuel resources of Griswold, L.S., 1892, Whetstones and the novaculites the southern part of the Oklahoma coal field. Part 4 – the of Arkansas: Arkansas Geological Survey Annual Re- Howe-Wilburton district: U.S. Geological Survey Bul- port for 1890, v. 3, 443p. letin 874-D, p. 255-298.

Gross, J.S., S.A. Thompson, B.L. Claxton, and M.B. Hendricks, T.A., L.S. Gardner, M.M. Knechtel, and Carr, 1995, Reservoir distribution and exploration po- P. Averitt, 1947, Geology of the western part of the tential of the Spiro sandstone in the Choctaw trend, Ouachita Mountains in Oklahoma: U.S. Geological Arkoma Basin, Oklahoma and Arkansas: American Survey Oil and Gas Investigations Series, Preliminary Association of Petroleum Geologists Bulletin, v. 79, p. Map 66, scale 1:42,240. 159-182. Hendricks, T.A., M.M. Knechtel, and J. Bridge, 1937, Ham, W.E., 1959, Correlation of pre-Stanley strata in Geology of Black Knob Ridge, Oklahoma: Bulletin of the Arbuckle – Ouachita regions, in L.M. Cline, W.J. the American Association of Petroleum Geologists, v. Hilseweck, and D.E. Feray, eds., The geology of the 21, p. 1-29. Ouachita Mountains, a symposium: Dallas Geological Society and Ardmore Geological Society, p. 71-86. Honess, C.W., 1923, Geology of the southern Ouachita Mountains of Oklahoma: Oklahoma Geological Survey Harlton, B.H., 1934, Carboniferous stratigraphy of the Bulletin 32, 278p. Ouachitas with special study of the Bendian: Bulletin of the American Association of Petroleum Geologists, v. Honess, C.W., 1930, Atoka, Pushmataha, McCurtain, 18, p. 1018-1049. Bryan, and Choctaw Counties, in Oil and gas in Okla-

32 FALL 2016 homa, v. 3: Oklahoma Geological Survey Bulletin 40, nary reports), 1917; Part 1, Metals and nonmetals ex- p. 83-108. cept fuels: U.S. Geological Survey Bulletin 660-C, p. C59-C122. Jordan, L., 1957, Subsurface stratigraphic names of Oklahoma: Oklahoma Geological Survey Guidebook 6, Miser, H.D., 1920, Mississippian tuff in the Ouachita 220p. Mountain region (abs.): Geological Society of America Bulletin, v. 31, p. 125-126. Knechtel, M.M., 1949, Geology and coal and natural gas resources of northern Le Flore County, Oklahoma: Miser, H.D., 1921, Llanoria, the Paleozoic land area in Oklahoma Geological Survey Bulletin 68, 76p. Louisiana and eastern : American Journal of Sci- ence, 5th series, v. 2, no. 8, p. 6-89. Laudon, R.B., 1958, Chesterian and Morrowan rocks in the McAlester Basin of Oklahoma: Oklahoma Geologi- Miser, H.D., 1926, Geologic map of Oklahoma: U.S. cal Survey Circular 46, 30p. Geological Survey and Oklahoma Geological Survey, scale 1:500,000. Laudon, R.B., 1959, Stratigraphy and zonation of the Stanley Shale of the Ouachita Mountains of Oklahoma: Miser, H.D., 1929, Structure of the Ouachita Mountains unpub. Ph.D. dissertation, University of Wisconsin, of Oklahoma and Arkansas: Oklahoma Geological Sur- Madison, Wisconsin, 117p. vey Bulletin 50, 30p.

Loomis, J., B. Weaver, and H. Blatt, 1994, Geochem- Miser, H.D., 1934, Carboniferous rocks of Ouachita istry of Mississippian tuffs from the Ouachita Moun- Mountains: Bulletin of the American Association of Pe- tains, and implications for the tectonics of the Ouachita troleum Geologists, v. 18, p. 971-1009. region, Oklahoma and Arkansas: Geological Society of America Bulletin, v. 106, p. 1158-1171. Miser, H.D., and C.W. Honess, 1927, Age relations of the Carboniferous rocks of the Ouachita Mountains of Lumsden, D.N., E.D. Pittman, and R.S. Buchanan, Oklahoma and Arkansas: Oklahoma Geological Survey 1971, Sedimentation and petrology of Spiro and Foster Bulletin 44, 28p. sands (Pennsylvanian), McAlester Basin, Oklahoma: American Association of Petroleum Geologists Bulle- Miser, H.D., and A.H. Purdue, 1929, Geology of the tin, v. 55, p. 254-266. DeQueen and Caddo Gap quadrangles, Arkansas: U.S. Geological Survey Bulletin 808, 195p. Mankin, C.J., and C.G. Dodd, 1963, Proposed refer- ence illite from the Ouachita Mountains of southeastern Montgomery, S., 1996, Jackfork Group: Ouachita Oklahoma: in A. Swineford, ed., Clays and clay miner- Mountains: Petroleum Frontiers, v. 13, no. 1, 59p. als, v. 10, Proceedings of the Tenth National Confer- ence on Clays and Clay Minerals: Pergamon Press, New Morris, R.C., 1971, Stratigraphy and sedimentology of York, p. 372-379. Jackfork Group, Arkansas: American Association of Pe- troleum Geologists Bulletin, v. 55, p. 387-402. Maravich, M.D., 1955, Developments in Oklahoma in 1954: Bulletin of the American Association of Petro- Morris, R.C., 1974, Carboniferous rocks of the Ouachi- leum Geologists, v. 39, p. 891-902. ta Mountains, Arkansas: A study of facies patterns along the unstable slope and axis of a flysch trough, in G. Marcher, M.V., and D.L. Bergman, 1983, Reconnais- Briggs, ed., Carboniferous of the southeastern United sance of the water resources of the McAlester and Tex- States: Geological Society of America Special Paper arkana quadrangles, southeastern Oklahoma: Oklahoma 148, p. 241-279. Geological Survey Hydrologic Atlas HA-9, sheet 1, scale 1:250,000. Mose, D., 1969, The age of the Hatton Tuff of the Ouachita Mountains, southeastern Oklahoma: Geologi- Miser, H.D., 1918, Manganese deposits of the Caddo cal Society of America Bulletin, v. 80, p. 2373-2378. Gap and DeQueen quadrangles, Arkansas, in Contribu- tions to economic geology (short papers and prelimi- NACSN, 1983, North American stratigraphic code:

33 OKLAHOMA GEOLOGICAL SURVEY

American Association of Petroleum Geologists Bulletin, tion? A summary: Oklahoma Geology Notes, v. 47, p. v. 67, p. 841-875. 204-209.

NACSN, 2005, North American stratigraphic code: Suneson, N.H., 1988, The geology of the Ti Valley Fault American Association of Petroleum Geologists Bulletin, in the Oklahoma Ouachita Mountains, in K.S. Johnson, v. 89, p. 1547-1591. ed., Shelf-to-basin geology and resources of Pennsylva- nian strata in the Arkoma Basin and Ouachita Mountains Niem, A.R., 1977, Mississippian pyroclastic flow and of Oklahoma: Oklahoma Geological Survey Guidebook ash-fall deposits in the deep-marine Ouachita flysch 25, p. 33-47. basin, Oklahoma and Arkansas: Geological Society of American Bulletin, v. 88, p. 49-61. Suneson, N.H., 1991, Re-evaluation of Jackfork Group type section along Indian Nation Turnpike, Oklahoma: Pitt, W.D., 1955, Geology of the core of the Ouachita Oklahoma Geology Notes, v. 51, p. 120-127. Mountains of Oklahoma: Oklahoma Geological Survey Circular 34, 34p. Suneson, N.H., and J.A. Campbell, 1990, Stop 8 – Pi- netop section, in N.H. Suneson, J.A. Campbell, and M.J. Pitt, W.D., R.O. Fay, L.R. Wilson, and J.A. Curiale, Tilford, eds., Geology and resources of the frontal belt of 1982, Geology of Pushmataha County, Oklahoma: East- the western Ouachita Mountains, Oklahoma: Oklahoma ern New Mexico University Studies in Natural Sciences Geological Survey Special Publication 90-1, p. 50-53. Special Publication No. 2, 101p. Suneson, N.H., I. Çemen, D.R. Kerr, M.T. Roberts, R.M. Pritchett, B.N., 2015, Map of Oklahoma oil and gas fields Slatt, and C.G. Stone, 2005, Stratigraphic and structural distinguished by coalbed methane and field boundaries: evolution of the Ouachita Mountains and Arkoma Basin, Oklahoma Geological Survey Geologic Map GM-40, southeastern Oklahoma and west-central Arkansas: Ap- scale 1:500,000. plications to petroleum exploration: Oklahoma Geologi- cal Survey Guidebook 34, 128p. Purdue, A.H., 1909, The slates of Arkansas: Arkansas Geological Survey, 95p. Suneson, N.H., and C.A. Ferguson, 1990, Geologic map of the Talihina 7.5’ quadrangle, Latimer and Le Flore Repetski, J.E., and R.L. Ethington, 1977, Conodonts Counties, Oklahoma: Oklahoma Geological Survey Map from graptolite facies in the Ouachita Mountains, Arkan- OGQ-7, scale 1:24,000. sas and Oklahoma, in C.G. Stone, ed., Symposium on the geology of the Ouachita Mountains – volume 1: Arkan- Suneson, N.H., and L.A. Hemish, eds., 1994a, Geology sas Geological Commission Miscellaneous Publication and resources of the eastern Ouachita Mountains frontal MP-13, p. 92-106. belt and southeastern Arkoma Basin, Oklahoma: Okla- homa Geological Survey Guidebook 29, 294p. Seely, D.R., 1963, Structure and stratigraphy of the Rich Mountain area, Oklahoma and Arkansas: Oklahoma Suneson, N.H., and L.A. Hemish, 1994b, Stop 19 – Geological Survey Bulletin 101, 173p. Johns Valley Shale at Stapp, in N.H. Suneson and L.A. Hemish, eds., 1994, Geology and resources of the east- Shaulis, B.J., T.J. Lapen, J.F. Casey, and D.R. Reid, 2012, ern Ouachita Mountains frontal belt and southeastern Ar- Timing and rates of flysch sedimentation in the Stanley koma Basin, Oklahoma: Oklahoma Geological Survey Group, Ouachita Mountains, Oklahoma and Arkansas, Guidebook 29, p. 103-106. U.S.A.: Constraints from U-Pb zircon ages of subaque- ous ash-flow tuffs: Journal of Sedimentary Research, v. Taff, J.A., 1901, Coalgate folio, Indian Territory: U.S. 82, p. 833-840. Geological Survey Atlas of the United States Folio GF- 74, 6p. Shelburne, O.B., Jr., 1960, Geology of the Boktukola Syncline, southeastern Oklahoma: Oklahoma Geological Taff, J.A., 1902, Description of the Atoka quadrangle Survey Bulletin 88, 84p. [Indian Territory]: U.S. Geological Survey Geologic At- las, Folio 79, 8p. Suneson, N.H., 1987, Lynn Mountain or Atoka Forma-

34 FALL 2016 Taff, J.A., and G.I. Adams, 1900, Geology of the east- troleum Geologists, Tulsa, Oklahoma, 140p. ern Choctaw coal fields, Indian Territory: U.S. Geologi- cal Survey, 21st Annual Report, part 2, p. 263-311. Whiteside, J.R., and R.C. Grayson, Jr., 1990, Carbonif- erous faunas, northern Ouachita Mountains, Tomlinson, C.W., 1929, The Pennsylvanian System in Oklahoma, in N.H. Suneson, J.A. Campbell, and M.J. the Ardmore Basin: Oklahoma Geological Survey Bul- Tilford, eds., Geology and resources of the frontal belt letin 46, 79p. of the western Ouachita Mountains, Oklahoma: Okla- homa Geological Survey Special Publication 90-1, p. Tulsa Geological Society, 1947, Field conference in 149-167. western part of the Ouachita Mountains in Oklahoma: Guidebook, Tulsa Geological Society, 56p. Wilmarth, M.G., 1938, Lexicon of geologic names of the United States: U.S. Geological Survey Bulletin 896, Ulrich, E.O., 1911, Revision of the Paleozoic systems: 2396p. Geological Society of America Bulletin, v. 22, p. 281- 680. Wilson, C.W., Jr., 1935, Age and correlation of Pennsyl- vanian surface formations, and of oil and gas sands of Ulrich, E.O., 1927, Fossiliferous boulders in the Ouachi- Muskogee County, Oklahoma: Bulletin of the American ta “Caney” Shale and the age of the shale containing Association of Petroleum Geologists, v. 19, p. 503-520. them: Oklahoma Geological Survey Bulletin 45, 48p. Wilson, C.W., Jr., and N.D. Newell, 1937, Geology of Visher, G., C. Stone, and B. Haley, 1978, Field guide the Muskogee – Porum district, Muskogee and McIn- to structure and stratigraphy of the Ouachita Mountains tosh Counties, Oklahoma: Oklahoma Geological Sur- and the Arkoma Basin for 1978 annual meeting of the vey Bulletin 57, 184p. AAPG, Oklahoma City: American Association of Pe-

About the Author

Neil Suneson started working with the OGS in 1986, mapping the northern Ouachita Mountains and Arkoma Basin as part of the STATEMAP project. Following this work, he and his colleagues completed some reconnaissance mapping in northwest Oklahoma and detailed mapping of the Oklahoma City metro area. Before joining the Survey, Neil received his Ph.D. from the University of California Santa Barbara, after which he worked for Chevron Resources Company and Chevron USA in the Bay Area. In addition to his work on Oklahoma stratigraphy and petroleum plays, Neil has taught the University of Oklahoma’s field camp in Colorado and a course on subsurface methods in petroleum exploration and development. He has published a number of papers and guidebooks, mostly on the stratigraphy and structure of southern Oklahoma. His current interest is completing “Roadside Geology of Oklahoma” for Mountain Press Publishing Company. 35 OKLAHOMA GEOLOGICAL SURVEY

Looking Down the Road Coming up next in The Oklahoma Geology Notes

2015 Rockfall By David Brown

In the Winter 2016 issue of the Oklahoma Geology Notes, OGS Geologist David Brown looks back at the 2015 rockfall on Inter- state 35. Other states like Colorado have far more frequent and substantial landslides that impact people more seriously, so this oc- currence was a bit of a surprise, and something the OGS will need to look at more carefully. Human alteration of the landscape can create unstable land surfaces just about anywhere.

36